1 Temperature control as key factor for optimal biohydrogen

2 production from thermomechanical pulping wastewater

3

4

5 Paolo Dessì a,* , Estefania Porca b, Aino–Maija Lakaniemi a, Gavin Collins b, Piet N. L. Lens a,c

6

7 a Tampere University of Technology, Faculty of Natural Sciences, P.O. Box 541, FI-33101

8 Tampere, Finland

9 bMicrobial Communities Laboratory, School of Natural Sciences, National University of Ireland

10 Galway, University Road, Galway, H91 TK33, Ireland

11 cUNESCO–IHE, Institute for Water Education, Westvest 7, 2611AX Delft, The Netherlands

12

13 Manuscript submitted to: Biochemical Engineering Journal

14

15 *Corresponding author: Phone: +39 3897651830, e-mail: [email protected], mail: Tampere

16 University of Technology, P.O. Box 541, FI-33101 Tampere, Finland

17 Estefania Porca: [email protected]

18 Aino-Maija Lakaniemi: [email protected]

19 Gavin Collins: [email protected]

20 Piet N.L. Lens: [email protected]

21

22 1

23 Abstract

24 This study evaluates the use of non-pretreated thermo-mechanical pulping (TMP) wastewater as a

25 potential substrate for hydrogen production by dark fermentation. Batch incubations were

26 conducted in a temperature gradient incubator at temperatures ranging from 37 to 80 °C, using an

27 inoculum from a thermophilic, xylose-fed, hydrogen-producing fluidised bed reactor. The aim was

28 to assess the short-term response of the microbial communities to the different temperatures with

29 respect to both hydrogen yield and composition of the active microbial community. High

30 throughput sequencing (MiSeq) of the reversely transcribed 16S rRNA showed that

31 Thermoanaerobacterium sp. dominated the active microbial community at 70 °C, resulting in the

-1 32 highest hydrogen yield of 3.6 (± 0.1) mmol H 2 g COD tot supplied. Lower hydrogen yields were

33 obtained at the temperature range from 37 to 65 °C, likely due to consumption of the produced

34 hydrogen by homoacetogenesis. No hydrogen production was detected at temperatures above 70

35 °C. Thermomechanical pulping wastewaters are released at high temperatures (50 to 80 °C), and

36 thus dark fermentation at 70 °C could be sustained using the heat produced by the pulp and paper

37 plant itself without any requirement for external heating.

38

39 Keywords

40 Dark fermentation, MiSeq, Pulp and paper mill wastewater, Thermoanaerobacterium ,

41 Thermomechanical pulping, Thermophilic

42

43

44

45

46

47

48 2

49 1. Introduction

50 Pulp and paper industry is facing an economic challenge due to globalised competition and

51 decreasing paper demand (Machani et al., 2014). The long-term success of the industry is believed

52 to be strictly linked to the ability of companies to innovate and create new value streams, which are

53 predicted to generate 40% of the companies’ turnover in 2030 (Toppinen et al., 2017). A biorefinery

54 concept, in which waste from the pulp and paper making process is used as a resource to generate

55 value-added products such as biofuels and biochemicals, is a promising strategy to expand the

56 product platform, reduce waste disposal costs and fulfil the environmental regulations on waste

57 emissions (Kinnunen et al., 2015; Machani et al., 2014; Moncada B. et al., 2016).

58

59 Pulping is the major source of polluted wastewaters of the whole papermaking process (Pokhrel and

60 Viraraghavan, 2004). Pulp mill wastewater is typically treated by the traditional activated sludge

61 process, but anaerobic processes have the advantages of coupling wastewater treatment to

62 renewable energy production, produce a lower quantity of waste sludge and require a smaller

63 volume than aerobic processes (Ashrafi et al., 2015). Among pulping processes, thermomechanical

64 pulping (TMP) produces a wastewater more suited for anaerobic biological processes than

65 chemical-based pulping, due to the low concentrations of inhibitory compounds such as sulphate,

66 sulphite, hydrogen peroxide, resin acid and fatty acids (Ekstrand et al., 2013; Rintala and Puhakka,

67 1994).

68

69 Thermomechanical pulping wastewater has been successfully used as a substrate for both

70 mesophilic (Gao et al., 2016) and thermophilic (Rintala and Lepistö, 1992) methane production via

71 anaerobic digestion. However, hydrogen (H 2) is a carbon free fuel expected to play a pivotal role in

72 energy production in the future (Boodhun et al., 2017). Dark fermentative H2 production has the

73 potential for energy recovery from waste paper hydrolysate (Eker and Sarp, 2017), pulp and paper

74 mill effluent hydrolysates (Lakshmidevi and Muthukumar, 2010) and even from untreated pulps 3

75 (Nissilä et al., 2012). Dark fermentative H2 production has also been reported from carbohydrate-

76 containing wastewaters, such as starch wastewater and palm oil mill effluent (Badiei et al., 2011;

77 Xie et al., 2014). Although TMP wastewaters are characterized by a high content of carbohydrates

78 (25 to 40% of the total COD) (Rintala and Puhakka, 1994), to our knowledge it has not yet been

79 tested as a substrate for H2 dark fermentation.

80

81 Thermophilic dark fermentation of TMP wastewater could be advantageous, as both biological

82 polysaccharide hydrolysis (Elsharnouby et al., 2013) and H2 yielding reactions (Verhaart et al.,

83 2010) are favoured by high temperature. High temperature also limits the growth of

84 homoacetogenic and methanogenic archaea (Oh et al., 2003), which may consume the

85 produced H2 in mixed culture systems. The main drawback of thermophilic processes is the energy

86 required to heat the reactors, but TMP wastewaters are released from the pulping process at a

87 temperature of 50 to 80 °C (Rintala and Lepistö, 1992), and could therefore be treated in

88 thermophilic bioreactors with minimal, or even without external heating.

89

90 Temperature is a key factor in dark fermentation, as even a change of a few degrees may result in

91 the development of a different microbial community and thus, affect the H2 yield (Dessì et al.,

92 2018; Karadag and Puhakka, 2010). Understanding of the composition of the microbial community

93 is also crucial in order to optimize the complex microbial H2 production process, involving both

94 hydrolytic and fermentative microorganisms (Kumar et al., 2017). Microbial communities from

95 dark fermentation of lignocellulose-based waste and wastewaters have been previously studied at

96 DNA level (Nissilä et al., 2012; Xie et al., 2014), but a RNA-based approach can provide more

97 detailed information on the microorganisms that produce (and consume) H2. Furthermore, the time

98 response on RNA changes is much faster than on DNA changes (De Vrieze et al., 2016), allowing

99 to detect the response of the microbial community to an environmental change in a relatively short

100 time. 4

101

102 In a previous study, a mixed culture was successfully adapted to thermophilic (70 °C) dark

103 fermentation of xylose in a fluidised bed reactor (FBR) and the H2 producing

104 Thermoanaerobacterium sp. accounted for > 99% of the active microbial community (Dessì et al.,

105 2018). In this study, the same adapted mixed culture was used to test if TMP wastewater is a

106 suitable substrate for dark fermentative H 2 production at various temperatures (from 37 to 80 °C),

107 and to describe how the active microbial community responds to the different temperatures.

108

109 2. Materials and methods

110 2.1 Source of microorganisms

111 The inoculum used in this study was biofilm-coated activated carbon originating from a

112 thermophilic fluidised bed reactor (FBR) used to study H2 production from xylose via dark

113 fermentation by gradually increasing the temperature of the reactor from 55 to 70 °C (Dessì et al.,

114 2018). The FBR was initially inoculated with heat-treated (90 °C, 15 min) activated sludge

115 originating from a municipal wastewater treatment plant (Viinikanlahti, Tampere, Finland). The

116 biofilm-coated activated carbon granules were sampled after 185 days of reactor operation, at that

117 point the FBR had been operated at 70 °C for 27 days. No xylose was present in the FBR medium at

118 the sampling time. The granules were stored at 4 °C for one week prior to utilisation. This inoculum

119 was used because the microbial community was dominated by Thermoanaerobacterium sp. (Dessì

120 et al., 2018), which previously showed potential for hydrolysis of lignocellulosic substrates and H2

121 production from the resulting sugars (Cao et al., 2014).

122

123 2.2 Wastewater characterization

124 The wastewater was collected from a pulp and paper mill located in Finland. It was the effluent of a

125 TMP process, in which wood was exposed to a high-temperature (120 °C) steam in order to obtain

126 the pulp. The wastewater had a temperature of about 70 °C at the time of the sampling, but was 5

127 cooled down and stored at 4 °C to minimise biological activity that might affect its composition.

128 The wastewater had a pH of 5.0 and a composition as given in Table 1.

129

130 Table 1 here

131

132 2.3 Temperature-gradient batch set–up

133 The batch cultures were conducted in anaerobic tubes with a total volume of 26 mL (17 mL

134 working volume and 9 mL headspace). The tubes were inoculated by adding 2 mL of biofilm-

135 coated activated carbon granules to 15 mL of TMP wastewater (Table 1). All the tubes were flushed

136 with N 2 for 5 min, and the internal pressure was equilibrated to atmospheric pressure by removing

137 the excess gas using a syringe and a needle before incubation. The initial pH of the batch cultures

138 (wastewater and inoculum) was adjusted to 6.3 (± 0.1) using 1 M NaOH, as higher pH may favour

139 the growth of methanogenic archaea (Jung-Yeol et al., 2012). The tubes were incubated at 200 rpm

140 shaking in a temperature-gradient incubator (Test Tube Oscillator, Terratec Asia Pacific, Australia ) at

141 37, 42, 48, 55, 59, 65, 70, 74 or 80°C (duplicate tubes at each temperature). The experiment was

142 interrupted after 111 hours, when no H 2 production was detected in any of the vials in two

143 consecutive samples, as long inactive periods may affect the RNA-level analysis (De Vrieze et al.,

144 2016).

145

146 Gas samples were collected for analysis 1 to 3 times per day. End-point liquid samples were

147 collected and stored at -20 °C before analysis. Non-inoculated control incubations, with fresh

148 activated carbon and TMP wastewater, were prepared at 37, 55 and 70 °C. Control incubations

149 containing 2 mL of fresh activated carbon and a mix of acetate and butyrate in Milli-Q® water (0.86

-1 150 g COD tot L each, 15 mL volume) were also prepared at 42, 65 and 80 °C to assess possible

151 adsorption of VFAs on virgin activated carbon.

152 6

153 2.4 Microbial community analyses

154 Biofilm-coated activated carbon granules and liquid medium were collected at the end of the

155 experiment and stored in 5 mL Eppendorf tubes at -80 °C. Microbial community analysis was

156 conducted separately on microbial communities growing attached to the granules and suspended in

157 the liquid medium, as the growth of suspended biomass was clearly visible in the vials after

158 incubation in the temperature range from 42 to 59 °C. Nucleic acids extraction using a modified

159 method from Griffiths et al. (2000), DNA inhibition, complementary DNA (cDNA) synthesis and

160 sequencing (using an Illumina MiSeq platform) were performed as described previously (Dessì et

161 al., 2018). Sequence analysis (1,395,864 sequences in total, 1,238,862 after quality check) was also

162 performed according to Dessì et al. (2018), but using a more recent version of Mothur (v1.39.5) and

163 Silva database (v128). The Illumina sequencing data was deposited to the NCBI Sequence Read

164 Archive under BioProject Number PRJNA428338.

165

166 2.5 Analytical methods

167 Gas production in the tubes was quantified by a volumetric syringe method (Owen et al., 1979), and

168 the gas composition was determined by gas chromatography-thermal conductivity detector (GC-

169 TCD) as reported previously by Dessì et al. (2017). Acetate, butyrate, ethanol, propionate, lactate,

170 and formate concentrations were measured with a high-performance liquid chromatograph (HPLC)

171 equipped with a refractive index detector (RID) (Shimadzu, Japan) and a Rezex RHM-

172 monosaccharide column (Phenomenex, USA) held at 40 °C. The mobile phase was 5 mM H 2SO 4

173 and the flow rate was 0.6 mL min -1. Glucose and xylose concentrations were measured using a

174 HPLC equipped with a RID and a RPM-monosaccharide column (Phenomenex, USA) held at 85 °C

175 with Milli-Q® water at a flow rate of 0.6 mL min -1 as the mobile phase. Furfural concentrations

176 were measured by gas chromatography-mass spectrometry (GC-MS) according to Doddapaneni et

177 al. (2018). Samples for HPLC and GC-MS analysis were filtered using 0.2 µm pore size filters.

178 Total chemical oxygen demand (COD tot ) and COD of the soluble compounds (COD s) was measured 7

179 using the dichromate method according to the Finnish standard SFS 5504. Initial and final pH of the

180 culture and the pH of the wastewater were determined using a WTW pH 330 meter equipped with a

181 Hamilton® Slimtrode probe (Sigma-Aldrich, USA). Total solids, volatile solids, total nitrogen and

3- 182 PO 4 -P were determined by the APHA standard procedures (APHA, 1998).

183

184 2.6 Calculations

185 Cumulative H 2 and CO 2 production was calculated according to Logan et al. (2002) and corrected

186 for temperature according to the Arrhenius equation. The theoretical COD tot was estimated from the

187 sum of the compounds detected by HPLC, according to the following equation (Van Haandel and

188 Van der Lubbe, 2012):

189

-1 190 COD tot = 8 ∙(4x+y-2z)/(12x+y+16z) g COD tot g CxHyOz (1)

191

192 where x, y and z are the number of C, H and O atoms in the organic molecule, respectively.

193

194 2.7 Statistical analysis

195 One-way analysis of variance (ANOVA) and the Tukey test (Box et al., 1978) at p = 0.05 were

196 conducted using the IBM SPSS Statistics package to assess significant differences in H2 yield after

197 incubation at different temperatures.

198

199 3. Results

200 3.1 Hydrogen production from TMP wastewater at the various temperatures

201 Batch incubations with TMP wastewater resulted in a different net H 2 yield at different

-1 202 temperatures (Figure 1; Table 2). The highest final H 2 yield of 3.6 (± 0.1) mmol H 2 g COD tot was

203 obtained in the batch cultures at 70 °C, in which H2 production started within 24 h of incubation and

204 remained stable after reaching the maximum (Figure 1). The maximum H 2 yield obtained at 65 °C 8

205 was comparable to the one obtained at 70 °C, but the produced H2 started to be consumed within 36

206 h resulting in a 51% lower final yield (Figure 1; Table 2). In the batch cultures at temperatures

207 lower than 70 °C, the H 2 produced was always partially (at 37, 42, 59 and 65 °C) or totally (at 48

208 and 55 °C) consumed. A negligible H2 production was obtained at both 74 and 80 °C (Figure 1), as

209 well as in the non-inoculated control incubations (Figure S1 in Supplementary Material).

210

211 Figure 1 here

212

213 Table 2 here

214

215 3.2 COD tot removal and metabolite production at the various temperatures

216 Similarly to H2 production yields, dark fermentation of TMP wastewater at the various temperatures

217 resulted in a different composition of the liquid phase (Figure 2). Acetate was the most abundant

218 metabolite detected in the temperature range from 37 to 70 °C. The final acetate concentration

-1 -1 219 increased with temperature from 0.34 (± 0.04) g CODtot L at 37 °C to 0.75 (± 0.18) g COD tot L at

-1 220 55 °C, and then decreased stepwise to 0.07 (± 0.00) and 0.08 (± 0.01) g COD tot L at 74 and 80 °C,

221 respectively (Figure 2). Butyrate was found regardless of the incubation temperature, with a final

-1 -1 222 concentration ranging from 0.06 (± 0.00) g COD tot L at 70 °C to 0.19 (± 0.00) g COD tot L at 59

- 223 °C. Ethanol was produced at 37, 42, 59, 65 and 70 °C, with a maximum of 0.14 (± 0.02) g COD tot L

224 1 at 65 °C (Figure 2). Dark fermentation of TMP wastewater caused a pH decrease from the initial

225 value of 6.3: the final pH ranged from 5.7 to 6.1 after incubation at 42, 48, 55, 59, 74 and 80 °C, but

226 was only 5.5 (± 0.1) after incubation at 37 °C, 5.2 (± 0.1) at 65 °C and 5.3 (± 0.0) at 70 °C (Figure

227 2).

228

229 Figure 2 here

230 9

231 In the batch incubations at various temperatures, the COD tot removal efficiency ranged from 69.4%

232 at 74 °C to 79.7% at 42 °C, resulting in a decrease from the initial concentration of 2.86 (± 0.00) g

-1 -1 233 COD tot L to a final concentration ranging from 0.58 (± 0.23) g COD tot L at 42 °C and 0.88 (±

-1 234 0.06) g COD tot L at 74 °C (Table 3). The COD tot removal efficiency was likely overestimated due

235 to the adsorption of VFAs on the activated carbon: in the adsorption experiment (Figure S2 in

236 Supplementary Material), up to 27% of the acetate and 90% of the butyrate was, in fact, adsorbed

237 on the fresh activated carbon after 111 h of incubation. The COD tot measured was comparable to the

238 COD tot estimated (using Eq. 1) by the sum of sugars and volatile fatty acids in the liquid phase after

239 incubation in the temperature range from 42 to 65 °C (Table 3) . However, the difference between

-1 240 measured and estimated COD tot was about 0.20 g CODtot L at 37, 70 and 80 °C, and even higher at

-1 241 74 °C (0.51 g COD tot L ).

242

243 Table 3 here

244

245 3.3 Effect of temperature on the active microbial community

246 Incubation temperature clearly affected the composition of the active microbial communities

247 growing for 111 h on TMP wastewater (Figure 3, Table 4). At 37 °C, Clostridium sp. accounted for

248 84 and 90% of the attached and suspended active microbial community, respectively. Higher

249 temperature resulted in a gradual decrease of the relative abundance of Clostridium sp., being 54%

250 of the attached active microbial community and < 2% of the suspended active microbial community

251 after incubation at 55 °C (Figure 3). Clostridium sp. was not detected either in the attached or

252 suspended active community after incubation at temperatures ≥ 59 °C (Figure 3). A bacterium

253 belonging to the order of Bacillales closely related to B. coagulans (Table 4) was detected in the

254 active attached and suspended microbial communities after incubation at 42 °C, with a relative

255 abundance of 14 and 10%, respectively, and only in suspended form after incubation at 48 °C, with

256 a relative abundance of 50% (Figure 3). 10

257

258 The relative abundance of Thermoanaerobacterium sp. (99% similarity to T.

259 thermosaccharolyticum ) among the attached active microorganisms gradually increased with

260 temperature, being only 2% after incubation at 37 °C and 87% at 59 °C (Figure 3, Table 4).

261 Thermoanaerobacterium sp. was also the most common suspended active microorganism after

262 incubation at 55 and 59 °C, with a relative abundance of 96 and 83%, respectively. After incubation

263 at 65 °C, the relative abundance of Thermoanaerobacterium sp. in the attached and suspended

264 active microbial community decreased to 57 and 25%, respectively, whereas unclassified

265 , with 92% similarity to Calditerricola sp. (Table 4) were found with a relative

266 abundance of 30 and 28%, respectively. After incubation at 70 °C, Thermoanaerobacterium sp. was

267 again the dominant active microorganism in both attached and suspended form, with a relative

268 abundance of 88 to 89%. After incubation at 59 and 70 °C, Caldanaerobius sp. was also found in

269 both attached and suspended form with relative abundance below 10% (Figure 3). After incubation

270 at both 74 and 80 °C, the RNA concentration was not high enough to perform the analysis due to

271 poor microbial growth, and thus microbial communities from 74 and 80 °C could not be analysed.

272

273 Figure 3 here

274

275 Table 4 here

276

277 4. Discussion

278 4.1 Fermentation of TMP wastewater at different temperatures

279 Hydrogen production from TMP wastewater inoculated with biofilm-coated activated carbon

280 granules was observed at a wide temperature range from 37 to 70 °C (Figure 1). The highest final

-1 -1 281 H2 yield of 3.6 (± 0.1) mmol H 2 g COD tot supplied, or 4.9 mmol H 2 g COD tot consumed, was

282 obtained at 70 °C (Table 2), which could be expected as the inoculum was collected from an FBR 11

283 operated at 70 °C (Dessì et al., 2018). The H2 yield obtained in this study is of the same order of

284 magnitude compared to previous studies on thermophilic direct dark fermentation of industrial,

-1 285 sugar-containing wastewaters. For example, Xie et al. (2014) obtained 5.8 mmol H 2 g COD tot from

286 starch wastewater at 55°C by a mixed culture dominated by T. thermosaccharolyticum , whereas

-1 287 Khongkliang et al. (2017) obtained 11.4 mmol H 2 g COD tot from starch wastewater by a pure

288 culture of T. thermosaccharolyticum .

289

290 The thermophilic active mixed microbial community previously enriched on xylose in the FBR was

291 dominated by microorganisms closely related to Thermoanaerobacterium thermosaccharolyticum

292 (Dessì et al., 2018). Changing of the substrate from xylose to TMP wastewater marginally impacted

293 the active microbial community in the temperature range 59 to 70 °C, as most of the sequences

294 obtained from the RNA samples matched T. thermosaccharolyticum (Table 4). A mixed culture

-1 295 dominated by T. thermosaccharolyticum has been shown to produce 7 mmol H 2 g cellulose at 70

296 °C (Gadow et al., 2013), showing potential for the one-step conversion of lignocellulosic materials

297 to H 2, avoiding a costly hydrolysis step. In fact, the Thermoanaerobacterium includes strains

298 of cellulolytic microorganisms, such as some strains of T. thermosaccharolyticum , able to hydrolyse

299 both cellulose and hemicellulose, and produce H2 from the resulting monosaccharides (Cao et al.,

300 2014). In this study, however, the microbial community analysis conducted at genus level does not

301 allow to assess possible cellulolytic capabilities of the detected Thermoanaerobacterium sp.

302

303 Although the inoculum was enriched for dark fermentation at 70 °C, H2 production at 70 °C

304 occurred only after 24 h of incubation (Figure 1). This is probably due to the handling of the

305 inoculum, which was stored at 4 °C for one week prior to being used for this experiment. Changes

306 in gene expression and DNA replication were shown to occur in Thermoanaerobacter

307 tengcongensis as response to a cold shock (Liu et al., 2014), as could be the case for the

308 Thermoanaerobacterium sp. dominating the active microbial community of the inoculum used in 12

309 this study. Although Thermoanaerobacterium sp. was the most abundant microorganism (relative

310 abundance close to 90%) in both the attached and suspended microbial community at both 59 and

311 70 °C, its relative abundance was lower at 65 °C (Figure 3). The same phenomenon was observed in

312 the FBR from where the inoculum originated (Dessì et al., 2018), and was attributed to either the

313 decreased activity of Thermoanaerobacterium sp. or to the increased activity of competing

314 microorganisms at 65 °C.

315

316 Despite the inoculum was enriched for thermophilic dark fermentation, H2 was already produced

-1 317 after 12 h of incubation at 37 °C, reaching a maximum yield of 3.2 (± 0.1) mmol H 2 g COD tot

-1 318 supplied within 24 h (Figure 1). A maximum yield of only 0.9 mmol H 2 g COD tot supplied was

319 previously obtained at 37 °C from a paper mill wastewater (production process type and wastewater

320 fraction used not specified) using heat treated digested sludge as inoculum (Marone et al., 2017).

321 The H 2 yields obtained in this study are also higher than those reported by Lucas et al. (2015) by

322 mesophilic (37 °C) dark fermentation of cassava, dairy and citrus wastewaters, which produced 1.4,

-1 323 1.7 and 1.3 mmol H 2 g COD tot supplied, respectively. This confirms the potential of TMP

324 wastewater for dark fermentation.

325

326 Clostridium sp. proliferated at 37 °C accounting for more than 80% of both the attached and

327 suspended active microbial community at the end of the batch incubation (Figure 3). It is plausible

328 that Clostridium sp. were present in the parent activated sludge but inactive in the FBR operated at

329 70 °C (Dessì et al., 2018). In fact, Clostridium spp. produce spores to survive under harsh

330 conditions, and are able to restore their metabolic activity after desporulation as soon as the

331 environmental conditions become more favourable (Li and Fang, 2007). Clostridium sp. cells might

332 also have been present in the TMP wastewater, which was not sterilised. However, the absence of

333 H2 and CO 2 in the non-inoculated control incubation at 37 °C (Figure S1 in Supplementary

334 Material) suggests that Clostridium sp. did not proliferate in the absence of the inoculum. 13

335

336 In this study, no H2 was produced at 74 or 80 °C (Figure 1) and the RNA concentration was too low

337 to allow sequencing analysis, suggesting a lack of active . This was attributed to the source

338 of inoculum used, as species within the Thermoanaerobacterium genus, such as T.

339 thermosaccharolyticum , may be inhibited by temperatures higher than 70 °C (Ren et al., 2008).

340 Gadow et al. (2013) obtained H 2 production from cellulose by a mixed microflora from a sewage

341 sludge digester even at 75 and 80 °C. However, H2 production at such high temperatures was

342 attributed to Thermoanaerobacter tengcongensins (Gadow et al., 2013), which was not part of the

343 active microbial community in this study. Some degradation products of hemicellulose such as

344 furfural or hydromethylfurfural may inhibit fermentative microorganisms (Jönsson et al., 2013),

345 including Thermoanaerobacterium , at a concentration over 1 g L -1 (Cao et al., 2010). However, the

346 TMP process is conducted at temperatures below 120 °C, which is too low to produce such high

347 concentrations of these inhibitory compounds (Baêta et al., 2017). In fact, the concentration of

348 furfural in the TMP wastewater used in this study was below the detection limit of the GC-MS

349 (Table 1).

350

351 A decrease in the cumulative H 2 production occurred in all the incubations at temperatures lower

352 than 70 °C (Figure 1), probably due to the activity of homoacetogenic bacteria. Homoacetogenesis,

353 in which 4 moles of H2 and 2 mol of CO 2 are consumed per mol of acetate produced, often occurs

354 in batch H 2 production experiments within the first 80 h of incubation, especially under mesophilic

355 conditions (for a review, see Saady, 2013). However, in this study, H2 seems to be consumed faster

356 under thermophilic (from 48 to 65 °C) as compared to mesophilic (37 °C) conditions (Figure 1),

357 suggesting that homoacetogenic microorganisms were mainly thermophiles or moderate

358 thermophiles. The CO 2 concentration in the batch incubations did not decrease as expected in case

359 of homoacetogenesis (Figure S3 in Supplementary Material). However, this could be explained

360 considering that CO 2 production may occur also through non-hydrogenic pathways, mainly the 14

361 ethanol production pathway (Figure 2). In the non-inoculated control incubations, CO 2 was also

362 detected, together with acetate, at both 55 and 70 °C, where H2 production was not observed (Figure

363 S1 in Supplementary Material). This suggests that non-hydrogenic, CO 2 producing pathways other

364 than ethanol production could have occurred as well.

365

366 Homoacetogens are among the most phylogenetically diverse functional groups of bacteria (Drake

367 et al., 2006). Among the thermophiles, Moorella thermoacetica , which accounted for 5% of the

368 suspended active community at 55 °C and 6% of the attached active community at 65 °C (Figure 3),

369 is a known homoacetogenic bacterium with an optimum growth temperature ranging from 55 to 60

370 °C (Drake et al., 2006). Clostridium spp. have also been previously found in thermophilic

371 fermentative reactors and associated with homoacetogenesis (Ryan et al., 2008). It is plausible that

372 the shift to autotrophic metabolism (e.g. homoacetogenesis) occurred after substrate depletion, as

373 suggested by Oh et al. (2003).

374

375 4.2 COD tot balance and metabolite production

376 The COD tot measured in the beginning of the incubations (Table 3) was 15% lower than the value

377 obtained while characterizing the TMP wastewater (Table 1). Apparently, some biological or non-

378 biological reaction occurred while storing the TMP wastewater at 4 °C before the experiment,

379 resulting in a slight COD tot concentration decrease. The COD tot removal efficiency during the

380 incubations was 69 to 80% regardless the incubation temperature (Table 3). It is in line with the

381 COD tot removal from anaerobic digestion of pulp and paper wastewater reported in the literature

382 (Meyer and Edwards, 2014), but higher than expected for dark fermentation which usually removes

383 only 30 to 40% of the COD tot (Sharma and Li, 2010). This was likely due to the adsorption of VFAs

384 on the activated carbon (Figure S2 in Supplementary Material), which caused an overestimation of

385 the COD tot removal. However, it should be noted that the adsorption experiment (Figure S2 in

386 Supplementary Material) was performed with fresh activated carbon, whereas the main experiment 15

387 was conducted with biofilm-covered activated carbon. The latter could have been partially saturated

388 with VFAs at the moment of inoculation, as VFAs were also produced in the FBR from where the

389 inoculum originated (Dessì et al., 2018).

390

391 In the temperature range from 42 to 65 °C, more than 85% of the residual COD tot was detected as

392 acetate, butyrate or ethanol by HPLC analysis (Table 3). However, 30 to 37% of the residual COD tot

393 was not detected as compounds identified by HPLC analysis after incubation at 37, 70 and 80 °C,

394 and even 58% of the residual COD tot was not identified after incubation at 74 °C. At 74 and 80 °C,

395 most of the undetected COD tot was likely constituted by polysaccharides such as cellulose, which

396 were not degraded due to the lack of bacterial activity at such high temperatures. At 74 and 80 °C,

397 CO 2 was also not produced (Figure S3 in Supplementary Material), supporting this conclusion.

398 Lignin, which accounts for 16-49% of the COD tot in TMP wastewater (Rintala and Puhakka, 1994)

399 can release VFAs at temperatures around 80 °C (Veluchamy and Kalamdhad, 2017), suggesting that

400 the acetate and butyrate detected at 74 and 80 °C (Figure 2) were produced physically rather than

401 biologically.

402

403 The simultaneous production of acetate and butyrate suggests that H2 was produced via both the

404 acetate and butyrate pathway in the temperature range from 37 to 70 °C. Acetate was the main

405 metabolite found in the liquid phase at all temperatures tested, excluding 74 and 80 °C (Figure 2),

406 and was associated either to H2 production through the acetate dark fermentative pathway or H2

407 consumption by homoacetogenesis. Interestingly, acetate production increased with temperature in

408 the range from 37 to 55 °C, and then decreased stepwise at temperatures above 55 °C (Figure 2). In

-1 -1 409 particular, the high (> 0.7 g COD tot L ) acetate (Figure 2) and concomitant low (< 0.5 mmol g

410 COD tot ) cumulative H2 yield (Figure 1) suggest that the optimum growth temperature for

411 homoacetogenic bacteria was about 55 °C in this study. At 70 °C, however, the H2 produced was

16

412 not consumed during the incubation (Figure 1), suggesting inhibition of homoacetogenic

413 microorganisms.

414

415 Solventogenesis occurred both in mesophilic (37 and 42 °C) and thermophilic (59, 65, and 70 °C)

416 batch cultures, resulting in ethanol production (Figure 2). Clostridium sp., which dominated the

417 active microbial communities under mesophilic conditions (Figure 3), may shift its metabolism

418 from acidogenesis to solventogenesis as response to a change of pH or volatile fatty acids

419 concentration, but the mechanism which triggers solventogenesis is not well understood (Kumar et

420 al., 2013). A pure culture of T. thermosaccharolyticum has been reported to produce ethanol

421 together with acetate and butyrate by dark fermentation of cellulose and complex lignocellulosic

422 substrates such as corn cob, corn straw and wheat straw (Cao et al., 2014). Similarly, in this study,

423 acetate, butyrate and ethanol were the main metabolites (Figure 2) of the dark fermentation of TMP

424 wastewater at 65 and 70 °C by a mixed culture dominated by T. thermosaccharolyticum (Figure 3;

425 Table 4).

426

427 4.3 Practical implications

428 Hydraulic retention times lower than 24 hours are typically used for dark fermentation of

429 wastewater in bioreactors operated in continuous mode (Lin et al., 2012). Therefore, based on the

430 results obtained in this batch experiment (Figure 1), dark fermentation of TMP wastewater at 37 and

431 65 °C appears favourable if suspended biomass bioreactors are used, as homoacetogenic bacteria

432 would be flushed out (Figure 1). However, bioreactors retaining high active biomass content, such

433 as FBRs or upflow anaerobic sludge bioreactors (UASBs), would enable higher organic loadings

434 and conversion rates than suspended biomass bioreactors (Koskinen et al., 2006). Therefore,

435 attached biomass bioreactors operated in continuous mode at 70 °C are recommended for H 2

436 production via dark fermentation of TMP wastewater. A proper insulation and temperature control

437 are nevertheless necessary to keep the temperature inside the bioreactor accurately at 70 °C, as a 17

438 decrease of 5 °C may already result in a decreased efficiency due to H2 consumption by

439 homoacetogenic bacteria. However, H2 production at 70 °C can be quickly restored in case of

440 failure of the temperature control. In fact, H2 production was detected at 70 °C within only 24 h

441 (Figure 1) with a thermophilic inoculum previously stored at 4 °C for one week.

442

443 Despite the surprisingly high COD tot removal efficiency of 69 to 80 % obtained in this study (Table

444 3), dark fermentation of TMP wastewater resulted in the generation of an effluent containing 0.5 to

-1 445 1.0 g COD tot L (Table 3), mainly in the form of VFAs, thus requiring further treatment prior

446 discharge to the environment. Such effluent can be either treated by a traditional activated sludge

447 plant, or further valorised by producing energy or high value chemicals. Promising strategies for the

448 valorisation of dark fermentation effluents include further H2 production by photofermentation or

449 microbial electrolysis cells, methane production by anaerobic digestion, and bioplastics or

450 electricity production using microbial fuel cells (for reviews, see Ghimire et al., 2015 and Bundhoo,

451 2017).

452

453 5. Conclusions

454 Hydrogen was produced by dark fermentation from TMP wastewater at a wide range of

455 temperatures (37 to 70 °C) using a mixed microbial community enriched on xylose at thermophilic

456 conditions. An operation temperature of 70 °C was the most favourable for dark fermentative H2

457 production and effectively repressed the activity of homoacetogenic bacteria. Therefore,

458 considering that TMP wastewater is produced at elevated temperature, dark fermentation at 70 °C

459 may be a cost-effective approach for the treatment and valorisation of this wastewater. However,

460 temperature must be efficiently controlled, as a shift of only a few degrees may decrease the H 2

461 yield.

462

463 Acknowledgements 18

464 This work was supported by the Marie Skłodowska-Curie European Joint Doctorate (EJD) in

465 Advanced Biological Waste-To-Energy Technologies (ABWET) funded from Horizon 2020 under

466 grant agreement no. 643071.

467

468 References

469 APHA, 1998. Standard Methods for the Examination of Water and Wastewater, twentieth ed.

470 American Public Health Association/American Water Works Association/Water Environment

471 Federation, Washington DC.

472 Ashrafi, O., Yerushalmi, L., Haghighat, F., 2015. Wastewater treatment in the pulp-and-paper

473 industry: A review of treatment processes and the associated greenhouse gas emission. J.

474 Environ. Manage. 158, 146–157.

475 Badiei, M., Jahim, J.M., Anuar, N., Abdullah, S.R.S., 2011. Effect of hydraulic retention time on

476 biohydrogen production from palm oil mill effluent in anaerobic sequencing batch reactor. Int.

477 J. Hydrogen Energy 36, 5912–5919.

478 Baêta, B.E.L., Cordeiro, P.H. de M., Passos, F., Gurgel, V.A.L., de Aquino, S.F., Fdz-Polanco, F.,

479 2017. Steam explosion pretreatment improved the biomethanization of coffee husks.

480 Bioresour. Technol. 245, 66–72.

481 Boodhun, B.S.F., Mudhoo, A., Kumar, G., Kim, S.-H., Lin, C.-Y., 2017. Research perspectives on

482 constraints, prospects and opportunities in biohydrogen production. Int. J. Hydrogen Energy

483 42, 27471–27481.

484 Box, G.E.P., Hunter, W.G., Hunter, J.S., 1978. Statistics for experimenters: An introduction to

485 design, data analysis, and model building. John Wiley and sons.

486 Bundhoo, Z.M.A., 2017. Coupling dark fermentation with biochemical or bioelectrochemical

487 systems for enhanced bio-energy production: A review. Int. J. Hydrogen Energy 42, 26667–

488 26686.

489 Cao, G.-L., Zhao, L., Wang, A.-J., Wang, Z.-Y., Ren, N.-Q., 2014. Single-step bioconversion of 19

490 lignocellulose to hydrogen using novel moderately thermophilic bacteria. Biotechnol. Biofuels

491 7, 82.

492 Cao, G., Ren, N., Wang, A., Guo, W., Xu, J., Liu, B., 2010. Effect of lignocellulose-derived

493 inhibitors on growth and hydrogen production by Thermoanaerobacterium

494 thermosaccharolyticum W16. Int. J. Hydrogen Energy 35, 13475–13480.

495 De Vrieze, J., Regueiro, L., Props, R., Vilchez-Vargas, R., Jáuregui, R., Pieper, D.H., Lema, J.M.,

496 Carballa, M., 2016. Presence does not imply activity: DNA and RNA patterns differ in

497 response to salt perturbation in anaerobic digestion. Biotechnol. Biofuels 9, 244.

498 Dessì, P., Lakaniemi, A.-M., Lens, P.N.L., 2017. Biohydrogen production from xylose by fresh and

499 digested activated sludge at 37, 55 and 70 °C. Water Res. 115, 120–129.

500 Dessì, P., Porca, E., Waters, N.R., Lakaniemi, A.-M., Collins, G., Lens, P.N.L., 2018. Thermophilic

501 versus mesophilic dark fermentation in xylose-fed fluidised bed reactors: Biohydrogen

502 production and active microbial community. Int. J. Hydrogen Energy 43, 5473–5485.

503 Doddapaneni, T.R.K.C., Jain, R., Praveenkumar, R., Rintala, J., Romar, H., Konttinen, J., 2018.

504 Adsorption of furfural from torrefaction condensate using torrefied biomass. Chem. Eng. J.

505 334, 558–568.

506 Drake, H.L., Küsel, K., Matthies, C., 2006. Acetogenic Prokaryotes, in: Springer (Ed.), The

507 Prokaryotes. New York, pp. 354–420.

508 Eker, S., Sarp, M., 2017. Hydrogen gas production from waste paper by dark fermentation: Effects

509 of initial substrate and biomass concentrations. Int. J. Hydrogen Energy 42, 2562–2568.

510 Ekstrand, E.-M., Larsson, M., Truong, X.-B., Cardell, L., Borgström, Y., Björn, A., Ejlertsson, J.,

511 Svensson, B.H., Nilsson, F., Karlsson, A., 2013. Methane potentials of the Swedish pulp and

512 paper industry – A screening of wastewater effluents. Appl. Energy 112, 507–517.

513 Elsharnouby, O., Hafez, H., Nakhla, G., El Naggar, M.H., 2013. A critical literature review on

514 biohydrogen production by pure cultures. Int. J. Hydrogen Energy 38, 4945–4966.

515 Gadow, S.I., Jiang, H., Hojo, T., Li, Y.-Y., 2013. Cellulosic hydrogen production and microbial 20

516 community characterization in hyper-thermophilic continuous bioreactor. Int. J. Hydrogen

517 Energy 38, 7259–7267.

518 Gao, W.J., Han, M.N., Xu, C.C., Liao, B.Q., Hong, Y., Cumin, J., Dagnew, M., 2016. Performance

519 of submerged anaerobic membrane bioreactor for thermomechanical pulping wastewater

520 treatment. J. Water Process Eng. 13, 70–78.

521 Ghimire, A., Frunzo, L., Pirozzi, F., Trably, E., Escudie, R., Lens, P.N.L., Esposito, G., 2015. A

522 review on dark fermentative biohydrogen production from organic biomass: Process

523 parameters and use of by-products. Appl. Energy 144, 73–95.

524 Griffiths, R.I., Whiteley, A.S., O’Donnell, A.G., Bailey, M.J., 2000. Rapid method for coextraction

525 of DNA and RNA from natural environments for analysis of ribosomal DNA- and rRNA-

526 based microbial community composition. Appl. Environ. Microbiol. 66, 5488–5491.

527 Jung-Yeol, L., Chen, X.-J., Lee, E.-J., Min, K.-S., 2012. Effects of pH and carbon sources on

528 biohydrogen production by co-culture of Clostridium butyricum and Rhodobacter sphaeroides .

529 J. Microbiol. Biotechnol. 22, 400–406.

530 Jönsson, L.J., Alriksson, B., Nilvebrant, N.-O., 2013. Bioconversion of lignocellulose: inhibitors

531 and detoxification. Biotechnol. Biofuels 6, 16.

532 Karadag, D., Puhakka, J.A., 2010. Effect of changing temperature on anaerobic hydrogen

533 production and microbial community composition in an open-mixed culture bioreactor. Int. J.

534 Hydrogen Energy 35, 10954–10959.

535 Khongkliang, P., Kongjan, P., Utarapichat, B., Reungsang, A., O-Thong, S., 2017. Continuous

536 hydrogen production from cassava starch processing wastewater by two-stage thermophilic

537 dark fermentation and microbial electrolysis. Int. J. Hydrogen Energy 42, 27584–27592.

538 Kinnunen, V., Ylä-Outinen, A., Rintala, J., 2015. Mesophilic anaerobic digestion of pulp and paper

539 industry biosludge–long-term reactor performance and effects of thermal pretreatment. Water

540 Res. 87, 105–111.

541 Koskinen, P.E.P., Kaksonen, A.H., Puhakka, J.A., 2006. The relationship between instability of H 2 21

542 production and compositions of bacterial communities within a dark fermentation fluidized-

543 bed bioreactor. Biotechnol. Bioeng. 97, 742–758.

544 Kumar, G., Sivagurunathan, P., Sen, B., Mudhoo, A., Davila-Vazquez, G., Wang, G., Kim, S.-H.,

545 2017. Research and development perspectives of lignocellulose-based biohydrogen production.

546 Int. Biodeterior. Biodegradation 119, 225–238.

547 Kumar, M., Gayen, K., Saini, S., 2013. Role of extracellular cues to trigger the metabolic phase

548 shifting from acidogenesis to solventogenesis in Clostridium acetobutylicum . Bioresour.

549 Technol. 138, 55–62.

550 Lakshmidevi, R., Muthukumar, K., 2010. Enzymatic saccharification and fermentation of paper and

551 pulp industry effluent for biohydrogen production. Int. J. Hydrogen Energy 35, 3389–3400.

552 Li, C., Fang, H.H.P., 2007. Fermentative hydrogen production from wastewater and solid wastes by

553 mixed cultures. Crit. Rev. Environ. Sci. Technol. 37, 1–39.

554 Lin, C.-Y., Lay, C.-H., Sen, B., Chu, C.-Y., Kumar, G., Chen, C.-C., Chang, J.-S., 2012.

555 Fermentative hydrogen production from wastewaters: A review and prognosis. Int. J.

556 Hydrogen Energy 37, 15632–15642.

557 Liu, B., Zhang, Y., Zhang, W., 2014. RNA-seq-based analysis of cold shock response in

558 Thermoanaerobacter tengcongensis , a bacterium harboring a single cold shock protein

559 encoding gene. PLoS One 9, 3.

560 Logan, B.E., Oh, S.-E., Kim, I.S., Van Ginkel, S., 2002. Biological hydrogen production measured

561 in batch anaerobic respirometers. Environ. Sci. Technol. 36, 2530–2535.

562 Lucas, S.D.M., Peixoto, G., Mockaitis, G., Zaiat, M., Gomes, S.D., 2015. Energy recovery from

563 agro-industrial wastewaters through biohydrogen production: Kinetic evaluation and

564 technological feasibility. Renew. Energy 75, 496–504.

565 Machani, M., Nourelfath, M., D’Amours, S., 2014. A mathematically-based framework for

566 evaluating the technical and economic potential of integrating bioenergy production within

567 pulp and paper mills. Biomass and Bioenergy 63, 126–139. 22

568 Marone, A., Ayala-Campos, O.R., Trably, E., Carmona-Martínez, A.A., Moscoviz, R., Latrille, E.,

569 Steyer, J.-P., Alcaraz-Gonzalez, V., Bernet, N., 2017. Coupling dark fermentation and

570 microbial electrolysis to enhance bio-hydrogen production from agro-industrial wastewaters

571 and by-products in a bio-refinery framework. Int. J. Hydrogen Energy 42, 1609–1621.

572 Meyer, T., Edwards, E.A., 2014. Anaerobic digestion of pulp and paper mill wastewater and sludge.

573 Water Res. 65, 321–349.

574 Moncada B., J., Aristizábal M., V., Cardona A., C.A., 2016. Design strategies for sustainable

575 biorefineries. Biochem. Eng. J. 116, 122–134.

576 Nissilä, M.E., Li, Y.-C., Wu, S.-Y., Lin, C.-Y., Puhakka, J.A., 2012. Hydrogenic and methanogenic

577 fermentation of birch and conifer pulps. Appl. Energy 100, 58–65.

578 Oh, S.-E., Van Ginkel, S., Logan, B.E., 2003. The relative effectiveness of pH control and heat

579 treatment for enhancing biohydrogen gas production. Environ. Sci. Technol. 37, 5186–90.

580 Owen, W.F., Stuckey, D.C., Healy Jr., J.B., Young, L.Y., McCarty, P.L., 1979. Bioassay for

581 monitoring biochemical methane potential and anaerobic toxicity. Water Res. 13, 485–492.

582 Pokhrel, D., Viraraghavan, T., 2004. Treatment of pulp and paper mill wastewater—A review. Sci.

583 Total Environ. 333, 37–58.

584 Ren, N., Cao, G., Wang, A., Lee, D., Guo, W., Zhu, Y., 2008. Dark fermentation of xylose and

585 glucose mix using isolated Thermoanaerobacterium thermosaccharolyticum W16. Int. J.

586 Hydrogen Energy 33, 6124–6132.

587 Rintala, J.A., Lepistö, S.S., 1992. Anaerobic treatment of thermomechanical pulping whitewater at

588 35-70°C. Water Res. 26, 1297–1305.

589 Rintala, J.A., Puhakka, J.A., 1994. Anaerobic treatment in pulp- and paper-mill waste management:

590 A review. Bioresour. Technol. 47, 1–18.

591 Ryan, P., Forbes, C., Colleran, E., 2008. Investigation of the diversity of homoacetogenic bacteria

592 in mesophilic and thermophilic anaerobic sludges using the formyltetrahydrofolate synthetase

593 gene. Water Sci. Technol. 57, 675–680. 23

594 Saady, N.M.C., 2013. Homoacetogenesis during hydrogen production by mixed cultures dark

595 fermentation: Unresolved challenge. Int. J. Hydrogen Energy 38, 13172–13191.

596 Sharma, Y., Li, B., 2010. Optimizing energy harvest in wastewater treatment by combining

597 anaerobic hydrogen producing biofermentor (HPB) and microbial fuel cell (MFC). Int. J.

598 Hydrogen Energy 35, 3789–3797.

599 Toppinen, A., Pätäri, S., Tuppura, A., Jantunen, A., 2017. The European pulp and paper industry in

600 transition to a bio-economy: A Delphi study. Futures 88, 1–14.

601 Van Haandel, A., Van der Lubbe, J., 2012. Handbook of biological wastewater treatment: Design

602 and Optimisation of Activated Sludge Systems. Quist Publishing, Leidschendam, The

603 Netherlands.

604 Veluchamy, C., Kalamdhad, A.S., 2017. Enhancement of hydrolysis of lignocellulose waste pulp

605 and paper mill sludge through different heating processes on thermal pretreatment. J. Clean.

606 Prod. 168, 219–226.

607 Verhaart, M.R.A., Bielen, A.A.M., Van der Oost, J., Stams, A.J.M., Kengen, S.W.M., 2010.

608 Hydrogen production by hyperthermophilic and extremely thermophilic bacteria and archaea:

609 Mechanisms for reductant disposal. Environ. Technol. 31, 993–1003.

610 Xie, L., Dong, N., Wang, L., Zhou, Q., 2014. Thermophilic hydrogen production from starch

611 wastewater using two-phase sequencing batch fermentation coupled with UASB methanogenic

612 effluent recycling. Int. J. Hydrogen Energy 39, 20942–20949.

613

614

615

616

24

617 Figures

618 Figure 1 – Hydrogen yield from batch incubation of thermomechanical pulping wastewater at

619 various temperatures (from 37 to 80 °C) using thermophilic biofilm-containing activated carbon as

620 inoculum. Error bars refer to the standard deviations of the duplicates.

621

622

623

624

625

626

627

628

629

25

630 Figure 2 – Concentration of detected sugars, volatile fatty acids and alcohols (on primary y-axis)

631 and pH (on secondary y-axis) after 111 h of incubation of thermomechanical pulping wastewater at

632 various temperatures (from 37 to 80 °C) using thermophilic biofilm-containing activated carbon as

633 inoculum. Error bars refer to the standard deviations of the duplicates.

634

635

636

637

638

639

640

641

642

643

644

26

645 Figure 3 – Relative abundance of the active genera resulting from MiSeq sequencing of the partial

646 16S rRNA (transcribed to 16S cDNA) on microbiological samples obtained from the biofilm-

647 containing activated carbon (attached) and from the liquid medium (suspended) after batch

648 incubation with thermomechanical pulping wastewater at various temperatures (from 37 to 70 °C).

649 The microbial genera are listed in order of relative abundance. Samples at 74 and 80 °C could not

650 be analysed due to the low RNA concentration present in the samples.

651

652

653

654

655

656 27

657 Table 1 - Composition of the thermomechanical pulping wastewater used in this study

Parameter Concentration

(mg L -1)

Total solids 3771 ± 10

Volatile solids 2452 ± 8

Total COD 3352 ± 82

Soluble COD 3289 ± 54

Total nitrogen < 10

3- Total PO 4 -P 2.8

Acetate < 30

Furfural < 10

Glucose 43 (± 2)

Xylose 38 (± 0)

658

659

660

661

662

663

664

665

666 28

667 Table 2 - Maximum and final hydrogen yield obtained from batch incubation of thermomechanical

668 pulping wastewater at various temperatures (from 37 to 80 °C) using thermophilic biofilm-

669 containing activated carbon as inoculum

-1 -1 Temperature H2 yield (mmol H 2 g COD tot H2 yield (mmol H 2 g Time required

(°C) supplied) COD tot consumed) for maximum

H2 yield (h)

Maximum Final Final

37 3.2 (± 0.1) 1.4 (± 0.1) 1.9 (± 0.2) 23

42 a 1.5 0.6 1.3 63

48 0.6 (± 0.1) 0.1 (± 0.0) 0.1 (± 0.0) 18

55 0.4 (± 0.1) 0.0 (± 0.0) 0.0 (± 0.0) 12

59 1.7 (± 0.8) 0.6 (± 0.3) 0.9 (± 0.5) 18

65 3.7 (± 0.4) 1.8 (± 0.2) 2.6 (± 0.3) 36

70 3.6 (± 0.1) 3.6 (± 0.1) 4.9 (± 0.4) 90

74 0.1 (± 0.0) 0.1 (± 0.0) 0.2 (± 0.0) n.a. b

80 0.0 (± 0.0) 0.0 (± 0.0) 0.0 (± 0.0) n.a.

670 a Hydrogen was produced only in one of the duplicate tubes;

671 b Not applicable.

672

673

674

675

676

677

678

29

679 Table 3 - COD tot balances after incubation of thermomechanical pulping wastewater at various

680 temperatures (from 37 to 80 °C) using thermophilic biofilm-containing activated carbon as

681 inoculum

Temperature Final COD tot Final COD tot Difference COD tot

(°C) measured a estimated b (measured – removal

(g L -1) (g L -1) estimated) (%) c

37 0.79 (± 0.00) 0.60 (± 0.04) 0.19 (± 0.04) 72.5

42 0.58 (± 0.23) 0.66 (± 0.12) -0.08 (± 0.11) 79.7

48 0.70 (± 0.01) 0.67 (± 0.00) 0.03 (± 0.02) 75.7

55 0.82 (± 0.14) 0.90 (± 0.22) -0.07 (± 0.08) 71.2

59 0.84 (± 0.03) 0.88 (± 0.01) -0.04 (± 0.04) 70.7

65 0.80 (± 0.04) 0.70 (± 0.03) 0.10 (± 0.00) 72.0

70 0.73 (± 0.10) 0.54 (± 0.03) 0.20 (± 0.07) 74.3

74 0.88 (± 0.06) 0.37 (± 0.00) 0.51 (± 0.07) 69.4

80 0.62 (± 0.06) 0.41 (± 0.02) 0.21 (± 0.05) 78.4

a -1 682 Data obtained by measurement according to the standard procedure; the initial COD tot was 2.86 g L ;

b 683 Data obtained by the sum of the COD tot equivalents (Eq. 1) of organic compounds measured in the liquid

684 phase;

c 685 Calculated from COD tot measured.

686

687

688

689

690

30

691 Table 4 - Association of the six most abundant 16S rRNA gene sequences to species collected in

692 the GenBank

Family Genus and species a Accession Matching Similarity

number sequence b (%) c

Thermoanaerobacteraceae Thermoanaerobacterium JX984971 474-765 99

thermosaccharolyticum

Clostridiaceae Clostridium sp. AY548785 450-741 99

Bacillaceae Bacillus coagulans MF373392 512-803 100

Bacillaceae Calditerricola NR_112684 529-820 92

yamamurae

Thermoanaerobacteraceae Caldanaerobius sp. LC127102 482-773 99

Thermoanaerobacteraceae Moorella thermoacetica CP017237 145404-145695 100

693 a Closest cultured species in GenBank;

694 b Section of the 16S rRNA gene (in bp) matching the sequence obtained by MiSeq analysis;

695 c Percentage of identical nucleotide pairs between the 16S rRNA gene sequence and the closest cultured

696 species in GenBank.

697

698

699

700

701

702

703

704 31

705 Supporting material

706 Figure S1 – Carbon dioxide yield profiles (a) and acetate yield after 111 h of incubation (b)

707 obtained in the non-inoculated incubation of thermomechanical pulping wastewater at 37, 55 and 70

708 °C. Hydrogen was not detected at any of the temperatures tested. Error bars refer to the standard

709 deviations of the duplicates.

710 Figure S2 – Adsorption of VFAs on activated carbon. Acetate and butyrate concentration before

711 and after 111 h of incubation with fresh activated carbon at 42, 65 and 80 °C. The initial

-1 712 concentration of VFAs was chosen hypothesizing that only 40% of the 2.86 g COD tot L was

-1 713 removed through dark fermentation, and equally distributing the remaining 1.71 g COD tot L

714 between acetate and butyrate. Error bars refer to the standard deviations of the duplicates.

715

716 Figure S3 – Carbon dioxide yield from batch incubation of thermomechanical pulping wastewater

717 at various temperatures (from 37 to 80 °C) using thermophilic biofilm-containing activated carbon

718 as inoculum. Error bars refer to the standard deviations of the duplicates.

32

Figure S1 – Carbon dioxide yield profiles (a) and acetate yield after 111 h of incubation (b) obtained in the non-inoculated incubation of thermomechanical pulping wastewater at 37, 55 and 70

°C. Hydrogen was not detected at any of the temperatures tested. Error bars refer to the standard deviations of the duplicates.

Figure S2 – Adsorption of VFAs on activated carbon. Acetate and butyrate concentration before and after 111 h of incubation with fresh activated carbon at 42, 65 and 80 °C. The initial

-1 concentration of VFAs was chosen hypothesizing that only 40% of the 2.86 g COD tot L was

-1 removed through dark fermentation, and equally distributing the remaining 1.71 g COD tot L between acetate and butyrate. Error bars refer to the standard deviations of the duplicates.

Figure S3 – Carbon dioxide yield from batch incubation of thermomechanical pulping wastewater at various temperatures (from 37 to 80 °C) using thermophilic biofilm-containing activated carbon as inoculum. Error bars refer to the standard deviations of the duplicates.

3.5 4.5 37 °C ) 4

added 3 42 °C 3.5 COD) COD 2.5 -1 348 °C g -1 2 g

2 2.5 2 55 °C 2 1.5 1.5 59 °C

1 yield (mmol H 1

2 65 °C H yield (mmol yield (mmol CO 0.5 2 0.5 70 °C

CO 0 0 0 74 °C 0 20 40 60 80 100 120 Time Time (hours) (hours) 80 °C